数理解析研究所講究録 1193 巻 2001 年 79-103 79

ASYMPTOTIC BEHAVIOR OF CONTINUOUS-TIME MARKOV CHAINS OF THE EHRENFEST TYPE WITH APPLICATIONS*

神奈川大学 工学部 成田清正 (Kiyomasa Narita) Faculty of Engineering, Kanagawa Univ.

Abstract Kac [6] considers the discrete-time Markov chain for the original Ehrenfest urn model and gives the expression of the $n$-step transition in terms of the Krawtchouk polyno- mials; moreover, Kac’s formulas for transition laws show that the model yields Newton’s law of cooling at the level of the evolution of the averages. Here we first give a version of Newton’s law of cooling for Dette’s [4] Ehrenfest urn model. We next discuss the continuous-time formulation of Markov chain for Krafft and Schaefer’s [9] Ehrenfest urn model and calculate law of transitions and first-passage times in both the reduced description with states counting the number of balls in the specified urn and the full description with states enumerating the vertices of the $N$-cube. Our results correspond to a generalization of the work of Bingham [3]. We also treat several applications of urn models to queueing network and reliability theory.

1. Introduction

Let $s$ and $t$ be two parameters such that $0

$(1- \frac{i}{N})s$ , if $j=i+1.$ ’

$\frac{i}{N}t$ , if $j=i-1$ , $p_{ij}=\{$ (1.1) $0<<=^{j}=^{N}$ $1-(1- \frac{i}{N})s-\frac{i}{N}t$ , if $j=i$ , ’

$0$ , otherwise,

which is proposed by Krafft and Schaefer [9]. In case $s=t=1$ , the Markov chain is the same as P. and T. Ehrenfest [5] use for illustration of the process of heat exchange between two bodies that are in contact and insulated from the outside. The temperatures are assumed to change in steps of one unit and are represented by the numbers of balls in two urns. The two urns are marked I and II and there are $N$ balls labeled 1, 2, . . . , $N$ and distributed in two urns. The chain is in state $i$ when there are $i$ balls in I. At each $1/N$ ball point of time $n$ one ball is chosen at random (i.e. with equal probabilities among

*This research was supported in part by $\mathrm{G}\mathrm{r}\mathrm{a}\mathrm{n}\mathrm{t}_{-}\mathrm{i}\mathrm{n}_{-\mathrm{A}\mathrm{i}}\mathrm{d}$ for General Scientific Research (No. 10640138), The Ministry of Education, Science, Sports and Culture. 80

numbers 1, 2, . . . , $N$ ) and moved from its urn to other urn by the following way. If it is in urn I, it is placed in II with probability $t$ and returned to I with probability l-t. If it

$\Pi$ is in , it is placed in I with probability $s$ and returned to II with probability l-s. The case where $s+t=1$ is the one-parameter Ehrenfest urn model as investigated by Karlin and $\mathrm{M}\mathrm{c}\mathrm{G}\mathrm{r}\mathrm{e}\mathrm{g}\mathrm{o}\mathrm{r}[7]$ . The case where $s=t$ is the thinning of the Ehrenfest urn model as proposed by Vincze [14] and Tak\’acs [13]. Remark 1.1. The process of the Markov chain in the model (1.1) corresponds to a $N$ ‘reduced description’ with $N+1$ states, $0,1,$ $\cdots,$ counting the number of balls in urn I. If we use a ‘full description’ with $2^{N}$ states, representing the possible configurations by $N$-tuples $i=$ $(i_{1}, \ldots, i_{k}, \ldots , i_{N})$ , where $i_{k}=1,0$ if ball $k$ is in urn I, II, respectively, one may identify the states with the vertices of the N-cube.

$a$ $b$ Let and denote real numbers such that $a,$ $b>-1$ or $a,$ $b<-N$ , and consider the following homogeneous Morkov chain with state space $\{0,1, \ldots , N\}$ and transition probabilities

$(1- \frac{i}{N})(\frac{a+1+i}{N+a+b+2})l\text{ノ}$, if $j=i+1$ ,

$\frac{i}{N}(\frac{N+b+1-i}{N+a+b+2})\nu$ , if $j=i-1$ , $p_{ij}=\{$ (1.2)

$1-(1- \frac{i}{N})(\frac{a+1+i}{N+a+b+2})$ $l \text{ノ}-\frac{i}{N}(\frac{N+b+1-i}{N+a+b+2})l^{\text{ノ}}$, if $j=i$ ,

$0$ , otherwise,

$\nu>0$ $0=^{pij=^{1}}<<$ $N$ where is arbitrary such that for all $i,j=0,1,$ $\ldots,$ . Remark 1.2. The two-parameter Ehrenfest urn model (1.1) is obtained by putting $a=su/(s+t),$ $b=tu/(s+t)$ and $\nu=s+t$ and taking the limit as $uarrow\infty$ . For the model (1.2), Dette [4] obtains the more general integral representations for the $n$-step transition probabilities and the (unique) stationary with respect to the spectral measure of random walk and shows the following remark. Remark 1.3. (i) The $n$ -step transition probabilities $p_{ik}^{(n)}$ in the model (1.1) are given by

$p_{ik}^{(n)}=( \frac{p}{q})^{k}\sum_{=x\mathrm{o}}^{N}p^{x}q^{N\prime}-x_{I\iota(X}i,p,$ $N)I\mathrm{i}_{k}(’)x,p,$$N(1- \frac{s+t}{N}X)^{n}$ , where $p=s/(s+t)$ and $q=t/(s+t)$ . Here $\{K_{n}(x,p, N) : n=0,1, \ldots, N\}$ is the family of Krawtchouk polynomials, which will appear in Remark 1.4 below. This result is also found by Krafft and Schaefer [9]. (ii) The (unique) stationary probability distribution $\pi=(\pi_{0}, \ldots , \pi_{N})$ in the model (1.2) is given by

$\pi_{j}=\frac{\beta(a+1+j,N+b+1-j)}{\beta(a+1,b+1)}$ , where $\beta(x, y)$ denotes the beta function, that is,

$\beta(x, y)=\frac{\Gamma(x)\Gamma(y)}{\Gamma(x+y)}=\int_{0}^{1}u^{x-1}(1-u)^{y-}1du$ for $x,$ $y>0$ . 81

For the model (1.1) with $s=t=1$ , Kac [6] finds the same form with (i) of Remark 1.3 and shows Newton’s law of cooling that concerns the estimate of the excess of the numbers of balls in urn I over the mean of the stationary probability distribution. Our purpose is the following: $\bullet$ For the model (1.2), we will show the same result with Newton’s law of cooling [Theorem 2.1]. $\bullet$ For the model (1.1), we will give the continuous-time formulation and obtain the law of transition probabilities and first-passage times in both the ‘reduced’ and the ‘full’ descriptions, using the method of Bingham [3] [Theorems 2.2-2.3]. Remark 1.4. With $0=0$ the Krawtchouk polynomials are defined by the hypergeometric series:

$K_{n}(x,p, N)$ $=$ $2F1(-n, -x;-N;1/p)$

$=$ $\sum_{\nu=0}^{n}(-1)^{\nu}\frac{(\begin{array}{l}n\nu\end{array})(\begin{array}{l}x\nu\end{array})}{(\begin{array}{l}N\nu\end{array})}(\frac{1}{p})^{\nu}$ $N$ , $n=0,1,$ $\ldots,$ .

When the meaning is clear, we write simply $K_{n}(x)$ instead of $K_{n}(x,p, N)$ . From Karlin

$\mathrm{M}_{\mathrm{C}}\mathrm{c}\mathrm{r}\mathrm{e}\mathrm{g}_{\mathrm{o}\mathrm{r}}[7]$ and we cite the properties for $I_{\mathrm{L}_{n}}’(X)$ , which will be used in after sections.

(i) $\sum_{n=0}^{N}I\mathrm{i}_{n}’(X)_{Z}n=(1+z)N-x(1-\frac{q}{p}Z)^{x}$ $N$ , $x=0,1,$ $\ldots,$ .

(ii) $I\mathrm{i}_{n}^{r}(0)=1$ , $I\mathrm{i}_{0}^{r}(x)=1$ .

$I\mathrm{i}_{n}^{\Gamma}(x, p, N)=I\mathrm{i}_{x}^{r}(n,p, N)$ $n,$ $N$ (iii) , $x=0,1,$ $\ldots,$ .

(iv) $I \mathrm{i}_{n}^{r}(N-x, p, N)=(-1)^{n}(\frac{q}{p})^{n}I\iota_{n}’(x, q, N)$ .

2. Theorems

We first concern ourselves with evaluation of the excess of the number of balls in urn I over the value averaged by the stationary probability distribution for the model (1.2). Theorem 2.1. Let $\{X_{n} : n=0,1, \ldots\}$ be the Markov chain with transition probabilities (1.2) with $a>-1$ and $b>-1$ . Let $\pi=(\pi_{0}, \ldots , \pi_{N})$ be the stationary probabilitity distribution given by (ii) of Remark 1.3, and denote $by<\pi>the$ mean of $\pi$ , that is, $< \pi>=\sum_{j0}^{N}=j\pi_{j}$ . Set $Y_{n}=X_{n}-<\pi>and$ put $e_{n}=E_{i}(Y_{n})$ , that is, the expected value of $Y_{n}$ given $X_{0}=i$ . Then we have the following:

(i) $< \pi>=N(\frac{a+1}{a+b+2})$ . 82

(ii) $e_{n}=(i-< \pi>)[1-\nu(\frac{a+b+2}{N(N+a+b+2)})]^{n}$ .

$T$ Remark 2.1. Suppose that the frequency of transitions is $\tau$ per second. Then in time there are $n=T\tau$ transitions. Write

$c=- \log[1-\nu(\frac{a+b+2}{N(N+a+b+2)})]^{\mathcal{T}}$

Then (ii) of the above theorem yields

$e_{n}=(i-<\pi>)\exp[-cT]$ . (2.1)

$b$ Let $a,$ and $\nu$ be taken as in Remark 1.2. Then

$< \pi>arrow N(\frac{s}{s+t})$ and l–v $( \frac{a+b+2}{N(N+a+b+2)})arrow 1-\frac{s+t}{N}$ in the limit as $uarrow\infty$ , which shows

$e_{n}= \{i-N(\frac{s}{s+t})\}\exp[-C\tau]$ with $c=- \log[1-\frac{s+t}{N}]^{\tau}$ (2.2)

Newton’s law of cooling in Kac [6] is the special case of (2.2) where $s=t=1$ . Remark 2.2. Consider the Markov chain $\{X_{n} : n=0,1, \ldots\}$ for the model (1.2) with $a=-N-1-l$ and $b=-N-1-k$ , where $k$ and $l$ are nonnegative integers. Then, as follows from Dette [4], the stationary probability distribution is the hypergeometric distribution:

$\pi_{j}=\frac{(\begin{array}{l}N+lj\end{array})(\begin{array}{l}N+kN-j\end{array})}{(\begin{array}{l}2N+\iota_{+}kN\end{array})}$ $N$ , $j=0,1,$ $\ldots,$ ,

$< \pi>=\sum_{j=0}^{N}j\pi_{j}=N(\frac{N+l}{2N+l+k})$ $X_{n}$ having its mean . For , define $e_{n}$ by the same way as that in Theorem 2.1. Then a straightforward calculation as taken in the proofs of (i) of Theorem 2.1 and Remark 2.1 leads us to the following result:

$e_{n}=E_{i}[x_{n}-< \pi>]=(i-<\pi>)[1-\nu(\frac{2N+l+k}{N(N+\iota+k)})]^{n}$

$e_{n}=(i--<\pi>)\exp[-cT]$ with $c=- \log[1-\nu(\frac{2N+l+k}{N(N+l+k)})]^{\tau}$

Namely, Theorem 2.1 and Remark 2.1 hold for the model (1.2) with the choice $a=-N-1-^{\iota}$ and $b=-N-1-k$ . In the following, let us consider the two-parameter Ehrenfest urn model (1.1). Then we will give the continuous-time formulation for this model. We first consider the reduced description in continuous-time. Let $X(u)$ be the number 83

of balls in urn I at time $u$ . We shall sometimes speak of this number as the state of the system. Then $X(u)$ is a random function which can take integer values from $0$ to $N$ . Associate with the $i\mathrm{t}\mathrm{h}$ ball a random function $X_{i}(u)$ defined as follows:

$i$ $X_{i}(u)=1$ if the th ball is in urn I at the time $u$ , (2.3) $X_{i}(u)=0$ otherwise.

Assumption 2.1. The family $\{X_{i}(u) : i=1,2, \ldots , N\}$ of random functions are inde- pendent. Each $X_{i}(u)$ is a Markov process with continuous-time parameter $u_{=}>0$ on the state space $\{0,1\}$ and governed by the following $Q$ -matrix (infinitesimal generator): $Q=$ . Remark 2.3. Under Assumption 2.1, the transition probability matrix $\exp[uQ]$ is given by

where

$p_{00}(u)=q+p\exp[-(S+t)u/N]$ , $p_{01}(u)=p-p\exp[-(s+t)u/N]$ , $p_{10}(u)=q-q\exp[-(s+t)u/N]$ , $p_{11}(u)=p+q\exp[-(S+t)u/N]$ , with parameters $00$ } is a homogeneous Markov chain with state space $\{0,1, \ldots, N\}$ and transition probabilities

$i,$ $P_{ik}(u)=P(X(u)=k|X(\mathrm{O})=i)$ , $k=0,1,$ $\ldots$ , N. (2.4)

Secondly we consider the continuous-time Markov chain in the full description. For this process write

$\overline{X}(u)=(X_{1}(u), x_{2}(u),$ $\ldots$ , $X_{N}(u))$ , where $X_{i}(u)$ is given by (2.3), and put

$\overline{P}_{ij}(u)=P(\tilde{X}(u)=j|\overline{X}(0)=i)$, $i,j\in V$, (2.5) where $V$ is the vertex-set. Then we obtain the next theorem. Theorem 2.2. Under Assumption 2.1, we have the following: (i) In the full $d_{eSC}\dot{n}pti_{\mathit{0}}nf$ the transition probabilities (2.5) are given by

$\overline{P}_{ij}(u)=\frac{1}{2^{N}}[p_{0}\mathrm{o}(u)+p_{11}(u)]^{N-m}$ $[$ pol $(u)+p_{1}\mathrm{o}(u)]^{m}$ , $m=|i-j|$ , 84

where $p\mathrm{o}0(u),$ $p_{0}1(u),$ $p1\mathrm{o}(u)$ and $p_{11}(u)$ are the functions as given in Remark 2.3, and for $N$-tuples $i=(i_{1}, \ldots, i_{k}, \ldots, i_{N})$ and $j=(j_{1}, \ldots,j_{k}, \ldots , j_{N})$ , the distance $|i-j|ofj$ from $i$ is defined by $|i-j|= \sum_{k=1}^{N}|i_{k}-j_{k}|$ . (ii) In the reduced $d_{esC}ripti_{on_{J}}$ the transition probabilities (2.4) are given by

$P_{ik}(u)=( \frac{p}{q})^{k}\sum_{=x0}^{N}p^{x}q^{N\prime}-x_{Ic_{i}}(x,p, N)\mathrm{A}_{k}(x,p, N)\exp[-X(S+t)u/N]$ , where $K_{n}(x,p, N)|$ are the Krawtchouk polynomials as given in Remark 1.4. Remark 2.4. Bingham [3] shows Theorem 2.2 for the particular case where $s=t=1$ (i.e. $p=q= \frac{1}{2}$ ). The form (i) of Remark 1.3 in the discrete-time parameter is a corollary of Theorem 2.2. Indeed, we condition on the total number of transitions in $[0, u]$ , which is Poisson with parameter $u$ :

$P_{ik}(u)= \sum_{=n0}\overline{p}ik(n)e-u\frac{u^{n}}{n!}\infty$ ,

where $\overline{p}_{ik}(n)$ stand for the $n$ -step transition probabilities in the discrete-time chain. Then, by (ii) of Theorem 2.2 we see

$\sum_{n=0}^{\infty}\overline{p}_{ik}(n)\frac{u^{n}}{n!}=P_{ik}(u)eu$

$=( \frac{p}{q})^{k}\sum_{x=0}^{N}p^{x}q^{N-x_{K_{i}(x,N}}p,)Kk(x,p, N)\exp[\{1-X(S+t)/N\}u]$ .

Equating the coefficients of $u^{n}/n!$ , we get the form of $\overline{p}_{ik}(n)$ which is just the same expression with (i) of Remark 1.3. Lastly, consider the Markov chain with transition probabilities (1.1) in the reduced description of the random walk on the $N$ -cube in discrete-time, starting at the origin (state $0$ ). By $\tau_{0N}$ denote the first-passage time to the opposite vertex (state $N$ ), and let

$F_{0N}(Z)$ and $\mu_{0N}$ be the probability generating function and the mean of $\tau_{0N}$ , respectively:

$F_{\mathit{0}N}(Z)=E[z^{T_{0N}}]$ and $\mu 0N=E[T_{0N}]$ .

$\overline{\tau}_{0N}$ $\overline{F}_{\mathit{0}N}(Z)$ Let be the continuous-time analogue of $\tau_{0N}$ . By and $\overline{\mu}_{0N}$ denote the Laplace- Stieltjes transform and mean of $\overline{\tau}_{0N}$ , repectively:

$\overline{F}_{\mathit{0}N}(z)=E[\exp(-Z\tilde{\tau}_{0}N)]$ and $\tilde{\mu}0N=E[\overline{T}_{0N}]$ .

Then we will investigate the asymptotics of the first passage to the opposite vertex as $N$ increases, extending the theorem of Bingham [3]. Theorem 2.3. Consider the model (1.1). Put

$p= \frac{s}{s+t}$ , $q= \frac{t}{s+t}$ and $\lambda=\frac{s+t}{N}$ .

In (iv), (v) and (vi) below, suppose that $s_{=}

(i) $F_{0N}(_{Z)}=[ \sum_{x=0}^{N}(-1)^{x}\{\frac{1}{1-z(1-\lambda X)}\}]$

$\cross[\sum_{x=0}^{N}(\frac{q}{p})^{x}\{\frac{1}{1-z(1-\lambda_{X})}\}]^{-1}$

(ii) $\overline{F}_{0N}(z)=F_{0}N(\frac{1}{1+z})=[\sum_{x=^{0}}^{N}(-1)^{x}\{\frac{1}{z+\lambda x}\}]$

$\mathrm{x}[\sum_{x=0}^{N}(\frac{q}{p})^{x}\{\frac{1}{z+\lambda x}\}]^{-1}$

(iii) $\mu \mathit{0}N=\overline{\mu}0N$ .

(iv) $\mu_{\mathit{0}N}=\frac{1}{\lambda}\sum_{=x1}^{N}[(\frac{q}{p})^{x}-(-1)^{x}]\frac{1}{x}$ .

(v) For large $N,$ $\mu_{0N}\sim\frac{1}{tp^{N}}$ .

(vi) $(tp^{N})\tau_{0}N$ converges in distribution to the exponential distribution with parameter 1 as $Narrow\infty$ .

3. Proof of Theorem 2.1

(i) Write $\square (z)$ for the probability generating function of the stationary probability distri- bution $\pi=$ $(\pi_{0}, \ldots , \pi_{N})$ that is given in (ii) of Remark 1.3:

$\square (z)=\sum_{j=0}^{N}\pi_{j^{Z}}j=\frac{1}{\beta(a+1,b\dagger 1)}\int_{0}^{1}u^{a}(1-u)b\{zu+(1-u)\}^{N}du$

with constants $a,$ $b>-1$ , where $\beta(x, y)=\Gamma(x)\Gamma(y)/\Gamma(x+y)$ .

From this we calculate the differential coefficient $\Pi’(1)$ and obtain the result. (ii) First of all we rewrite (1.2) as follows:

$p_{i,i+1}= \frac{\nu}{N(N+a+b+2)}\{N(a+1)+(N-a-1)i-i^{2}\}$ ,

$p_{i,i-1}= \frac{\iota \text{ノ}{N(N+a+b+2)}}\{(N+b+1)i-i^{2}\}$ ,

$p_{i,i}=1- \frac{\nu}{N(N+a+b+2\mathrm{I}}\{N(a+1)+(2N-a+b)i-2i^{2}\}$ . 86

So, a straightforward calculation yields

$(i+1)\cdot pi,i+1+(i-1)\cdot pi,i-1+i\cdot p_{i,i}$

$=i+ \frac{\nu}{N(N+a+b+2)}\{N(a+1)-(a+b+2)i\}$ . (3.1)

For the Markov chain with transition probabilities (1.2), let $E(X_{n}|X_{n-1})$ be the condi- tional expectation of $X_{n}$ given $X_{n-1}$ . Then (3.1) implies

$E(X_{n}|X_{n-1})=X_{n-1}+ \frac{\nu}{N(N+a+b+2)}\{N(a+1)-(a+b+2)X_{n-1}\}$ ,

and so

$E(X_{n}|X_{n-1})-x_{n-1}=- \nu(\frac{a+b+2}{N(N+a+b+2)})\{X_{n-1}-<\pi>\}$ (3.2)

with $< \pi>=N(\frac{a+1}{a+b+2})$ . For $Y_{n}=X_{n}-<\pi>$ , set $e_{n}=E_{i}(\mathrm{Y}_{n})$ , that is, the expected value of $Y_{n}$ given $X_{0}=i$ . Then, taking note of that $E_{i}(X_{n}-X_{n}-1)=E_{i}[E(X_{n}|X_{n-1})-$

$X_{n-1}]$ , by (3.2) we can calculate as follows:

$e_{n}=E_{i}(x_{n}-<\pi>)=E_{i}[(X_{n-1}-<\pi>)+(X_{n}-X_{n}-1)]$ $=E_{i}(X_{n-1}-<\pi>)+E_{i}(X-nxn-1)$

$=e_{n-1^{-\mathcal{U}}}( \frac{a+b+2}{N(N+a+b+2)})E_{i}(x_{n}-1-<\pi>)$

$=[1- \nu(\frac{a+b+2}{N(N+a+b+2)})]e_{n-1}$ .

Thus, by successive applications of the above relation, we get

$e_{n}=[1- \nu(\frac{a+b+2}{N(N+a+b+2)})]^{n}e_{0}$,

where $e_{0}=E_{i}(X_{0^{-}}<\pi>)=i-<\pi>_{\alpha}$ Hence the proof is complete.

4. Proof of Theorem 2.2

(i) Consider the matrix $\overline{P}(u)=(\overline{P}_{ij}(u))$ with $\overline{P}_{ij}(u)$ given by (2.5). Then we must verify

that $\overline{P}(u)$ is a stochastic matrix, satisfying $\overline{P}(u)arrow I$ as $uarrow \mathrm{O}$ , where $I$ is the identity matrix. By (2.3) and Remark 2.3 it is easy to see that the transition probabilities $\overline{P}_{ij}(u)$ can be expressed by the form of (i). Since for each $i$ there are we have

$m$ $\sum_{j}\overline{P}_{ij}(u)=\frac{1}{2^{N}}\sum_{m=0}^{N}[p\mathrm{o}\mathrm{o}(u)+p_{11}(u)]^{N-m}$ [pol $(u)+p_{1}0(u)$ ]

$= \frac{1}{2^{N}}[\{p_{0}0(u)+p11(u)\}+\{p01(u)+p10(u)\}]N$

$= \frac{1}{2^{N}}[\{p\mathrm{o}\mathrm{o}(u)+p01(u)\}+\{p1\mathrm{o}(u)+p11(u)\}]N$

$= \frac{1}{2^{N}}[\{q+p\}+\{q+p\}]^{N}=‘\frac{1}{2^{N}}[1+1]^{N}=1$ . 87

Namely, $\overline{P}(u)$ is stochastic. On the other hand, by Remark 2.3, since

$p_{0}\mathrm{o}(\mathrm{o})+p_{11}(0)=(q+p)+(p+q)=1+1=2$ and since

$p_{01}(\mathrm{o})+p_{1}\mathrm{o}(\mathrm{o})=(p-p)+(q-q)=0$, we have that $\overline{P}(u)arrow I$ as $uarrow \mathrm{O}$ . Further we note that each $\overline{P}_{ij}(u)$ has the right-hand derivative at $u=0$ , associating with the $Q$ -matrix $Q=(\overline{P}_{ij}’(0))$ such that

$\tilde{P}_{ij}’(0)=-\frac{s+t}{2}$ if $m=0$ , $\frac{s+t}{2N}$ if $m=1$ , $0$ if $m>1$ .

(ii) Consider the coordinate process $X_{i}(u)$ given by (2.3). Then we may define the prob- ability generating function of $X_{i}(u)$ by $P( X_{i}(u)=0)+zP( X_{i}(u)=1)$ , which is from Remark 2.3,

$p_{00}(u)+zp_{01}(u)=(q+pz)[1+( \frac{1-z}{q+pz})p\exp[-(S+t)u/N]]$ , or

$p_{10}(u)+zp_{11}(u)=(q+pz)[1-( \frac{1-z}{q+pz})q\exp[-(s+t)u/N]]$ ,

according as $X_{\mathrm{i}}(0)$ is $0$ or 1. Remember that $X(u)= \sum_{i=1}^{N}X_{i}(u)$ and $P_{ik}(u)=P(X(u)=$ $k|X(\mathrm{O})=i)$ . Now, suppose that there are initially $i$ balls in urn I and $N-i$ in urn II.

$X_{i}(u)$ $N$ Then, since the quantities , where $i=1,2,$ $\ldots,$ , are independent, the probability generating function of $X(u)$ is calculated as follows:

$\sum_{k=0}^{N}Pik(u)z^{k}=E_{i}[z^{X(u})]=\Pi_{j=1}^{N}E_{i}[z^{X_{\mathrm{J}}()}u]$ $(E_{i}(\cdot)=E(\cdot|X(\mathrm{O})=i))$

$=[p_{1}\mathrm{o}(u)+zp11(u)]^{i}[p_{0}\mathrm{o}(u)+Zp_{01}(u)]^{N-i}$

$=(q+pz)^{N}[1- \frac{q}{p}\xi]lN[1+\xi]-i$

$=(q+pz)N \sum_{x=0}^{N}K_{x}(i)\xi^{x}$ $( \xi=(\frac{1-z}{q+pz})p\exp[-(s+t)u/N])$

$= \sum_{x=0}^{N}K_{x}(i)pxqN-x\{[1+\eta]^{N-x}[1-\frac{q}{p}\eta]^{x}\}\exp[-x(S+t)u/N]$

$( \eta=\frac{p}{q}z)$

$= \sum_{x=0}^{N}I\mathrm{i}_{x}^{r}(i)pqxN-x\{\sum_{k=0}^{N}K_{k}(X)\eta^{k}\}\exp[-x(s+t)u/N]$

$= \sum_{k=0}^{N}(\frac{p}{q})^{k}[\sum_{x=0}^{N}p^{x}q^{N-x_{I}}\iota’(i)I\mathrm{i}_{k}^{\Gamma}(Xx)\exp[-X(s+t)u/N]]z^{k}$. 88

$K_{n}(x)$ $N$ Here with $n,$ $x=0,1,$ $\ldots$ , are the Krawtchouk polynomials appearing in Remark

1.4, and in the above equalities the property of the generating function of $I\mathrm{i}_{n}’(X)$ is used at twice. Note the symmetry relation such that $\mathrm{A}_{x}^{\nearrow}(i)=I1_{i}^{r}(x)$ . Then, equating the coefficients of $z^{k}$ , we obtain the expression of $P_{ik}(u)$ . Hence the proof is complete.

5. Proof of Theorem 2.3

(i) Consider the Markov chain with transition probabilities (1.1). Denote by $p_{ij}^{(n)}$ and $T_{ij}$ the $n$ -step transition probabilities and the first-passage times to the state $j$ for this chain starting at the state $i$ , respectively. Set

$P_{ij}(z)= \sum_{n=0}^{\infty}p^{()}ijz^{n}n$ and $F_{ij}(z)= \sum_{n=0}^{\infty}P(T_{ij}=n)z^{n}$ .

Then, since $P_{ij}(z)=P_{jj}(z)F_{ij}(\mathcal{Z})$ for $i\neq j$ , we see that $F_{0N}(Z)=P0N(Z)/P_{NN}(Z)$ . By Remark 1.3, $p_{ij}^{(n)}$ is expressed in terms of the Krawtchouk polynomials, which implies

$P_{0N}(z)= \sum_{=n0}^{\infty}p^{()}\mathrm{o}^{n}Nzn$

$= \sum_{n=0}^{\infty}(\frac{p}{q})^{N}\sum_{=x0}^{N}p^{x}q^{N-x}I\iota \mathrm{o}(x’)KN(X)[(1-\lambda x)z]^{n}$

$=( \frac{p}{q})^{N}\sum_{=x0}^{N}p^{x}q^{N-x}K0(X)I\{’N(x)[\frac{1}{1-z(1-\lambda x)}]$ .

By Remark 1.4, since $K_{0}(x)=1$ and $I \mathrm{t}^{r_{N}}(x)=I\mathrm{i}_{x}’(N)=(-1)^{x}(\frac{q}{p})^{x}$, we get

$P_{0N}(z)=p \sum_{x}^{N}N(-1=0)^{x}[\frac{1}{1-z(1-\lambda X)}]$ .

By the same way, we get

$P_{NN(Z)}= \sum_{=n0}^{\infty}p^{()}Nz^{n}=p^{N}\sum^{N}n_{N}x=0(\frac{q}{p})^{x}[\frac{1}{1-z(1-\lambda x)}]$ , from which follows (i). (ii) The time reversibility shows that $\overline{F}_{0N}(Z)=E(\exp[-z\overline{T}_{0}N])=F_{0N}(\frac{1}{1+z})$ . So, the above formula (i) with $z$ replaced by $1/(1+z)$ yields (ii). (iii) Evidently,

$\overline{\mu}0N=E(\overline{T}_{0N})=-\overline{F}_{0N}’(0)=-[\frac{d}{dz}F_{0N}(\frac{1}{1+z})]_{z=0}=F_{0N}’(1)=E(T_{0N})=\mu 0N$. 89

(iv) The above result (ii) can be rewritten by the form

$\overline{F}_{0N}(_{\mathcal{Z})}=(\frac{1}{z}+f_{1}(z))(\frac{1}{z}+f_{2}(_{Z}))-1$ ,

where $f_{1}(z)= \sum_{x=1}^{N}(-1)^{x(\begin{array}{l}Nx\end{array})}[\frac{1}{z+\lambda x}]$ and $f_{2}(z)= \sum_{x=1}^{N}(\frac{q}{p})^{x}[\frac{1}{z+\lambda x}]$ .

Since $f_{1}’(z)$ and $f_{2}’(z)$ exist at $z=0$ , and since

$- \overline{F}_{0N}’(\mathcal{Z})=-[\frac{g(z)-h(Z)}{(1+zf2(z))^{2}}]$ with functions

$g(z)=-f_{2}(z)+zf_{1}’(\mathcal{Z})+z^{2}f_{1}’(\mathcal{Z})f_{2(Z})$ and $h(z)=-f_{1}(z)+zf_{2}’(z)+z^{2}f_{1}(z)f’2(z)$ ,

we have that $\mu_{oN}=\overline{\mu}_{0N}=-\overline{F}_{0N}’(0)=f_{2}(0)-f_{1}(0)=\sum_{x=1}^{N}[(\frac{q}{p})^{x}-(-1)^{x}]\frac{1}{\lambda x}$ .

Here, the assumption on $s$ and $t$ such that $s

$( \frac{q}{p})^{x}-(-1)^{x}=(\frac{t}{s})^{x}-(-1)^{x}=^{0}>$ $N$ for $x=1,2,$ $\ldots$ , , which guarantees that $\mu_{0N=}0>$ , giving (iv). (v) Put

$a_{N}= \frac{1}{\lambda}\sum_{x=1}^{N}(\frac{q}{p})^{x}\frac{1}{x}$ and $b_{N}= \frac{1}{\lambda}\sum_{x=1}^{N}(-1)^{x}\frac{1}{x}$ . (5.1)

Then we evaluate $a_{N}-b_{N}$ for large $N$ . Set

$p(x)=( \frac{p^{N}}{1-p^{N}})(\frac{q}{p})^{x}$ $N$ , $x=1,2,$ $\ldots,$ , so that

$\sum_{x=1}^{N}p(x)=(\frac{p^{N}}{1-p^{N}})[(1+\frac{q}{p})N-1]=1$ .

Namely, $\{p(x) : x=1,2, \ldots, N\}$ is a probability distribution, having the probability generating function

$( \frac{p^{N}}{1-p^{N}})[(1+\frac{q}{p}Z)^{N}-1]$

and mean $Nq/(1-p^{N})$ . For large $N$ , the can be approximated by the standard normal distribution, and so the mass of $p(x)$ concentrates on the intervals of 90

the form $(Nq/(1-p^{N})-C\sqrt{N}, Nq/(1-p^{N})+C\sqrt{N})$ with a constant $C>0$ . Namely, on such intervals, $1/x\sim(1-p^{N})/Nq$ . Hence

$\sum_{x=1}^{N}p(x)(\frac{Nq}{1-p^{N}})\frac{1}{x}\sim 1$ $(Narrow\infty)$ , which implies

$\frac{1}{\lambda}\sum_{x=1}^{N}(\frac{q}{p})^{x}\frac{1}{x}\sim\frac{1}{\lambda}(\frac{1-p^{N}}{Nq})(\frac{1-p^{N}}{p^{N}})$

$=( \frac{N}{s+t})(\frac{s+t}{Nt})\frac{(1-p^{N})^{2}}{p^{N}}$ $(Narrow\infty)$ .

Here, $0

$a_{N}\sim\underline{1}$ $(Narrow\infty)$ . (5.2) $tp^{N}$

On the other hand, as Bingham [3, p.604] shows, we have

$\frac{N}{2}\sum_{x=1}^{N}(-1)^{x}\frac{1}{x}=o(2^{N})$ $(Narrow\infty)$ .

Notice that

$(tp^{N})b_{N}=(tp^{N})( \frac{2}{s+t})\{\frac{N}{2}\sum_{x=1}^{N}(-1)^{x}\frac{1}{x}\}$

$=(2p)^{N}( \frac{2t}{s+t})\frac{1}{2^{N}}\{\frac{N}{2}\sum_{x=1}^{N}(-1)^{x}\frac{1}{x}\}$

and consider that for $s_{=}

$b_{N}=o( \frac{1}{tp^{N}})$ $(Narrow\infty)$ . (5.3)

Therefore, (5.2) and (5.3) yield

$a_{N}-b_{N} \sim\frac{1}{tp^{N}}$ $(Narrow\infty)$ , giving (v). (vi) The above result (ii) shows

$E[ \exp\{-(tp^{N})\overline{T}_{0N^{Z}}\}]=\overline{F}_{0N}((tp^{N})_{Z})=\frac{B_{N}}{A_{N}}$ , (5.4) 91

where

$A_{N}= \frac{1}{(tp^{N})z}+\sum_{x=1}^{N}(\frac{q}{p})^{x}[\frac{1}{(tp^{N})_{\mathcal{Z}+\lambda X}}]\sim\frac{1}{(tp^{N})z}+a_{N}$ $(Narrow\infty)$

and

$B_{N}= \frac{1}{(tp^{N})z}+\sum_{x=1}^{N}(-1)^{x}[\frac{1}{(tp^{N})z+\lambda x}]\sim\frac{1}{(tp^{N})z}+b_{N}$ $(Narrow\infty)$

with $a_{N}$ and $b_{N}$ as defined by (5.1). Accordingly, combining (5.4) with (5.3) and (5.2), we have

$E[\exp\{-(tp)N\overline{\tau}0Nz\}]$ $=$ $\frac{\frac{1}{(tp^{N})z}+o(\frac{1}{tp^{N}})}{\frac{1}{(tp^{N})z}+\frac{1}{tp^{N}}+o(\frac{1}{tp^{N}})}$

$arrow$ $\frac{\frac{1}{z}}{\frac{1}{z}+1}=\frac{1}{1+z}$ $(Narrow\infty)$ .

Since a $X$ having the exponential distribution with parameter 1 satisfies that $E[\exp\{-zx\}]=1/(1+z)$ , the above relation implies that $(tp^{N})\tilde{\tau}_{0}N$ converges in distribution to the exponential distribution with parameter 1 as $Narrow\infty$ , and (vi) follows similarly or by (ii). Hence the proof is complete.

5. Complements

In Theorem 2.2 there is a limit distribution in each case: Since $\lim_{uarrow\infty}p_{\mathrm{o}0}(u)=\lim_{uarrow\infty}p_{1}0(u)=q=t/(s+t)$ and $\lim_{uarrow\infty}p_{01}(u)=\lim_{uarrow\infty^{p_{11}}}(u)=$ $p=s/(s+t)$ , we have

$\lim_{uarrow\infty}\overline{P_{ij}}(u)=\frac{1}{2^{N}}[q+p]N-m[p+q]^{m}=\frac{1}{2^{N}}$ , so that the limit law is uniform for the ‘full description’. On the other hand, by Remark 1.4, since $K_{i}(0)=I1_{k}^{r}(0)=1$ , we have

$\lim_{uarrow\infty}P_{ik}(u)=(\frac{p}{q})^{k}p^{0}q^{N0}-=p^{k}qN-k$ $i,$ $N$ , where $k=0,1,$ $\ldots,$ .

Namely, the limit law is binomial for the $\zeta \mathrm{r}\mathrm{e}\mathrm{d}\mathrm{u}\mathrm{c}\mathrm{e}\mathrm{d}$ description’.

By contrast, in Kac’s theorem [6] for the $n$ -step transition probability of the discrete- time Markov chain that is the Ehrenfest model (1.1) with $s=t=1,$ $\pi_{k}=2^{-N}$ gives an invariant (stationary) probability distribution, but there is no limit distribution: Indeed, let $\overline{\mathrm{p}}=(\overline{p_{ik}})$ be the transition probability matrix for the model (1.1) with $s=t=1$ . Then it is easy to see that

$\lim_{narrow\infty}\overline{\mathrm{p}}^{2n}\neq\lim\overline{\mathrm{p}}^{2n+1}narrow\infty$ . 92

If we start at any vertex that we relabel as the ‘origin’ $(0, \ldots, 0)$ , the vertex it takes longest to hit is $(1, \ldots, 1)$ . Now if we apply a unit voltage between these two vertices so that the voltage at $(0, \ldots, 0)$ is 1 and the voltage at $(1, \ldots, 1)$ is $0$ , then all vertices having the same number of l’s share the same voltage and can be shorted. In general what we obtain is a new graph with $N+1$ vertices, where the $k$ th new vertex consists of the shorting of all vertices in the unit cube with $k1’ \mathrm{s}$ . Thus, as follows from Palacios [10] and [11], the $N$-cube yields the platonic graphs in electric networks. The well-known Ornstein-Uhlenbeck process is obtained in the diffusion limit of the Ehrenfest model (1.1) with $s=t=1$ , where we centre at $N/2$ , scale space and time and let $Narrow\infty$ . The Ehrenfest urn models are applicable to the field of neuron firing and have a comple- mentary aspect which is the study of the time taken to approach stationarity-measured in terms of variation distance-and the ‘cut-off phenomenon’ typically exhibited. Our work for a generalization is inspired by the above point of view.

6. Discrete-time Ehrenfest urn models with applications

The transition matrix $P=(p_{ij})$ as defined in (1.1) or (1.2) is a special cae of tridiagonal matrix $P=$ (6.1) satisfying

$\lambda_{i}=pi,i+1$ , $\mu_{i}=pi,i-1$ , $\nu_{i}=p_{i,i}$ $(i=0,1, \ldots, N)$ , $\lambda_{N}=0$ , $\mu_{0}=0$ .

This chain is a random walk on $\{0,1, \ldots, N\}$ . The model (1.2) is a generalization of the model (1.1) such that the transition probabilities $p_{i,i+1}$ and $p_{i,i-1}$ are allowed to be quadratic functions of the current state $i$ . In this section, for simplicity of the notation and the calculation, we take the discrete-time Ehrenfest model (1.1), that is, the model (6.1) satisfying

$\lambda_{i}=(1-\frac{i}{N})s$ , $\mu_{i}=\frac{i}{N}t$ , $\nu_{i}=1-(\lambda_{i}+\mu_{i})$ , (6.2)

$N$ where $0

one-parameter Ehrenfest model, we interpret $s$ and $t(=1-s)$ as probabilities of choosing urns I and II, respectively. If $s\neq 1$ and $t\neq 1$ , then the chain is a randam walk with sticky boundaries at $0$ and $N$ . 93

6.1 Steady state and time reversibility

The Ehrenfest model (6.1) with rates (6.2) satisfies the global balance equations

$\pi_{i}=\sum_{j=0}^{N}\pi jp_{ji}$ $N$ for $i=0,1,$ $\ldots,$

and also satisfies the detailed balance equations

$\pi_{i}p_{i}j=\pi_{j}pji$ $N$ for $i,j=0,1,$ $\ldots,$ .

Hence the chain is time reversible, having (unique) stationary probability distribution

$\pi_{j}=(\frac{s}{s+t})^{j}(\frac{t}{s+t})^{N-j}$ $N$ , $j=0,1,$ $\ldots$ , , where $0

6.2 Convergence $\mathrm{v}\mathrm{s}$ . recurrence

Here we discusss the Ehrenfest model (6.1) satisfying (6.2) with $s=t=1$ . Then we can show that recurrence is not observable for states far from the steady state $M=N/2$ , $N$ assuming that is even. For instance, the average time to reach $0$ from state $M$ , where $M$ is the mean for the binomial distribution (stationary probability distribution), is

$\frac{1}{2M}2^{2M}(1+O(\frac{1}{M}))$

whereas the average time to reach state $M$ from state $0$ is less than

$M+M\log M+o(1)$ .

With $M=10^{4}$ balls and rate of transition one ball per second, the return time to equi- librium from state $0$ is on the order of 102103 seconds, which is less than 29 hours (only $10^{6}000$ about a day), whereas it would take on the order of years to go from $M$ to $0$ , which is an astronomical time (see Bhattacharya and Waymire [2, pp.250-251]). Newton’s law of cooling (Theorem 2.1) is another manifestation of thermodynamic irreversibility in the Ehrenfest model. On the other hand, in much more general settings, several authors investigate ‘relaxation time’, measuring the time taken for the random walk to approximate the limit law in variation norm (see Aldous [1] and Kijima [8]). Example 6.1 (Finite Buffer). A discrete-time communication channel with a finite buffer of size $N$ behaves as follows: During the $i\mathrm{t}\mathrm{h}$ slot, either a new message arrives (if there is room for it), with positive probability, or one of the messages in the buffer (if any) is trans- mitted, with positive probability, or there is no change. This system can be modelled by a Markov chain which is in state $i$ when there are $i$ messages in the buffer $(i=0,1, \ldots, N)$ .

$i$ From state , there are transitions to states $i-1$ (except when $i=0$), $i$ and $i+1$ (except when $i=N$), with probabilities

$\mu_{i}=_{Pi,i}-1$ , $\nu_{i}=p_{i,i}=1-(\lambda_{i}+\mu_{i})$ and $\lambda_{i}=p_{i,i+1}$ , 94

respectively. The chain with rates of the form (6.2) is an example of such a transmission. Example $6.2(Portfoli_{\mathit{0})}$ . There will be one risky investment opportunity (such as a common stock) whose market price changes with time in probabilistic way. This risky asset will form a Markov chain called a random walk with sticky boundaries. Prices $0$ and $N$ are (certainly or uncertainly) reflecting states. If the price now is $0$ , then at the next time instant the price will be 1 with certainty or uncertainty. Similarly, if the price is $N$ now, the price at the next instant of time will be $N-1$ or $N$ . Otherwise, for states between $0$ and $N$ , the price will increase by 1 with probability $\lambda_{i}=p_{i,i+1}$ , or decrease by 1 with probability $\mu_{i}=p_{i,i-1}$ , or remain with probability $\nu_{i}=p_{i,i}=1-(\lambda_{i}+\mu_{i})$ . The chain with rates of the form (6.2) is an example of such a market. We follow the progress of an investor who owns shares in this asset, and who also owns shares in a risk- free asset (such as a government bond) whose value is multiplied by a constant $\gamma>1$ , at each discrete instant of time. The investor can allocate his wealth between the risky and the risk-free asset at each instant, and also, if he chooses, consume some of his wealth. The investor is to decide how to change his porfolio and how to consume money as time progresses, in order to maximize his expected total consumption. In a finite time horizon investment-consumption problem, all wealth is consumed at the terminal time. In an infinite-horizon discounted problem, the invester will maximize the expected total discounted consumption. The optimal-value function for the control problem contains two parameters $s$ and $t$ such that $0

7. Continuou-time Ehrenfest urn models with applications

Here we treat the Markov process which is the continuous-time formulation of the Markov chain with transition matrix (6.1) satisfying (6.2). First of all, we note that the Krawtchouk polynomials $K_{n}(x,p, N)$ as described in Remark 1.4, which play an essential role in the continuous-time formulation, have the following equivalent formula:

$Ii_{n}^{r}(x,p, N)=C^{-1} \sum_{\nu=0}(-1)n\nu(\frac{q}{p})^{\nu}$ $N$ , $n=0,1,$ $\ldots$ ? , where $C=$ and $0

The Krawtchouk polynomials are orthogonal with respect to the binomial distribution $p^{x}q^{N-x}$ $N$ with masses at the $N+1$ points $x=0,1,$ $\ldots$ ; .

$N$ $k_{ij}$ For $i=0,1,$ $\ldots,$ , define by

$k_{ij}=I\mathrm{i}_{i}(\prime j,p, N)$ .

Then, as follows from Karlin and $\mathrm{M}\mathrm{c}\mathrm{G}\mathrm{r}\mathrm{e}\mathrm{g}\mathrm{o}\mathrm{r}[7]$ , the polynomials $k_{ij}$ satisfy the following recurrence relation:

$-jk_{ij}=iqk_{i-1,j}-[iq+(N-i)p]k_{ij}+(N-i)pk_{i}+1,j$ . 95

7.1 Expression of the Krawtchouk polynomials The above-mentioned formula and Remark 1.4 yield the following explicit expression of the polynomials $k_{ij}$ :

$k_{0j}=1$ , $k_{1j}=\underline{1}$ (Np-j), $pN$

$k_{Nj}=(-1)^{j}( \frac{q}{p})^{j}$ , $k_{N-1,j}=(-1)j \frac{q^{j-1}}{p’ N}$ (Nq-j),

$k_{ij}=k_{ji}$ $N$ , $i,j=0,1,$ $\ldots,$ .

Let $s$ and $t$ be the parameters such that $0

$p= \frac{s}{s+t’}$ $q= \frac{t}{s+t}$ ,

$K=(pk_{i}j)_{0}\leqq i,j\leqq N$ $\lambda_{j}=1-\frac{j}{N}(_{S}+t)$ $0<=^{j

Let $P$ be the transition matrix as given by (6.1) and (6.2). Then, it follows from Krafft and Schaefer [9] that $PK=K$ A. Namely, the Krawtchouk polynomials provide right

$P$ eigenvectors of corresponding to the eigenvalues $\lambda_{j}$ . 7.2 Q-matrix

Let $X(u)$ and $\overline{X}(u)$ be the Markov processes with continuos parameter $u\geq 0$ , which are the continuou-time formulation of the Ehrenfest urn model (1.1) in the ‘reduced description’ and the ‘full description’ as described in Remark 1.1, respectively. Then, the transition probabilities of $X(u)$ and $\overline{X}(u)$ are denoted by

$P_{ik}(u)$ $i,$ $N$ , where $k=0,1,$ $\ldots,$ and

$\overline{P}_{ij}(u)$ , where $i$ and $j$ vary in the vertex-set $V$, respectively (see (2.4) and (2.5)). By Theorem 2.2 and the expression of the Krawtchouk polynomials $\underline{k}ij$ in the preceding subsection, we can calculate the right-hand derivatives at $u=0$ for $P_{ij}(u)$ and $P_{ik}(u)$ , associating the $Q$ -matrices $Q=(\overline{P}_{ij}^{J}(0))$ and $Q=(P_{ik}’(0))$ with the following results:

$- \frac{1}{2}(s+t)$ if $m=|i-j|=0$ , (i) $Q=(\overline{P}_{ij(0}’))$ , $\overline{P}_{ij}’(\mathrm{o})=\{$ $\frac{1}{2N}(s+t)$ if $m=|i-j|=1$ , (7.1) $0$ if $m=|i-j|>1$ , where $i=(i_{1}, \ldots, i_{k}, \ldots, i_{N})$ and $j=(j_{1}, \ldots,j_{k}, \ldots,j_{N})$ are in the vertex-set $V$ , and $m=|i-j|= \sum_{k=1}^{N}|i_{k}-j_{k}|$ , and moreover $0

$\lambda_{i}$ if $k=i+1$ ,

$\mu_{i}$ if $k=i-1$ , (ii) $Q=(P_{ik}’(0))$ , $P_{ik}’(\mathrm{o})=\{$ (7.2) $-(\lambda_{i}+\mu_{i})$ if $k=i$ , $0$ otherwise, 96 where

$\lambda_{i}=(1-\frac{i}{N})s$ $\mu_{i}=\frac{i}{N}t$ $t<1=$ $N$ , , $0

The $Q$ -matrices (7.1) and (7.2) are called the (formal) infinitesimal generator associated with $\tilde{X}(u)$ and $X(u)$ , respectively, which represent the instantaneous transition rates. The Markov proxess $X(u)$ as governed by (7.2) is a finite birth-death process with state space $\{0,1, \ldots, N\}$ , satisfying the following Kolmogorov’s backward equation:

$\frac{d}{du}P_{i,k}(u)=(\frac{i}{N}t)Pi-1,k(u)-(\frac{i}{N}t+(1-\frac{i}{N})S)Pi,k(u)$

$+((1- \frac{i}{N})S)Pi+1,k(u)$ , (7.3)

$i,$ $N$ $P_{N+1,k}(u)\equiv 0$ $P_{-1,k}(u)\equiv 0$ $k=0,1,$ $\ldots,$ , , , and the initial conditions

$P_{i,k}(0)=1$ if $i=k$ , $P_{i,k}(0)=0$ if $i\neq k$ .

7.3 Models in 3-cube

(i) The Markov process $\overline{X}(u)$ in the full description. Define the vertex-set $V$ by

$V=\{\overline{0}, \overline{1}, \overline{2}, \overline{3}, \overline{4}, \overline{5}, \overline{6}, \overline{7}\}$ where

$\overline{0}=(0,0,0)$ , $\overline{1}=(0,0,1)$ , $\overline{2}=(0,1,0)$ , $\overline{3}=(0,1,1)$

$\overline{4}=(1,0,0)$ , $\overline{5}=(1,0,1)$ , $\overline{6}=(1,1,0)$ , $\overline{7}=(1,1,1)$ . Set

$c= \frac{1}{2}(s+t)$ and $d= \frac{1}{6}(s+t)$ .

Then, by (7.1), the $Q$ -matrix associated with $\overline{X}(u)$ on the vertex set $V$ has the following expression:

$-cd00d0d0-cd00d0d0-cd0d00d0-cdd00d00-cd0d0d00-cddd0000$ $-cddd0000$ $Q=($ $-cd000dd0$ ). 97

(ii) The Markov process $X(u)$ in the reduced description. By (7.2), the $Q$ -matrix associated with $X(u)$ on $\{0,1,2,3\}$ has the following expression:

$\frac{1}{3}t$ $- \frac{1}{3}(2s+t)$ $\frac{1}{3}2s$

$\backslash \text{ノ}-s00$ $- \frac{1}{3}(_{S}+2t0t)$ $Q=$ $\frac{1}{3}2t0s$ $\frac{1}{-3}0_{S}0t)$ .

Example 7.1(Mean First-Passage Time). Let us consider the Markov process $X(u)$ on the 3-cube in the reduced description, starting at the origin (state $0\Leftrightarrow\overline{0}=(0,0,0)$ ). Let $\overline{T}_{03}$ be the first-passage time of $X(u)$ to the opposite vertex (state $3\Leftrightarrow\overline{3}=(1,1,1)$ ), and set $\overline{\mu}_{03}=E[\overline{T}_{03}]$ . Then, by (iii) and (iv) of Theorem 2.3, we obtain

$\overline{\mu}_{03}=\frac{1}{s}\{9+(\frac{t}{s}-1)\frac{7}{2}+(\frac{t}{s})^{2}\}$ , and hence $\overline{\mu}_{03}=10$ if $s=t=1$ .

7.4 Reliability theory

$X_{i}(u)$ $N$ Let , where $i=0,1,$ $\ldots$ , , be the one-particle process as defined in (2.3), satisfying Assumption 2.1. Then we can interpret as follows: Each $X_{i}(u)$ sojourns in urn I (state 1) and urn II (state $0$ ) with exponentiol holding time of parameter $1/N$ , and at the next instant of the end of its sojourn time, $X_{i}(u)$ remains in urn I and urn II repeatedly with probabilities $1-t$ and l-s, respectively, or moves to the opposite urn with probabilities

$t$ and $s$ , respectively. The Markov process $X(u)$ in the reduced description is defined by

$X(u)= \sum^{N}X_{i}(ui=1)$ ,

with transition probabilities $P_{ik}(u)$ as given in (2.4). The probability law of $X(u)$ is obtained by the convolution of probability laws of i.i.d. random functions $X_{i}(u)$ . So, the rate of occurrence that $X(u)$ terminates its sojourn is estimated at $1=N*(1/N)$ . Namely, $X(u)$ is a kind of Markov renewal process such that the discrete-time Markov chain with transition probabilities (1.1) moves from state to state with exponential holding time of parameter 1. From this poin,of view, the results (ii) and (iii) of Theorem 2.3 are regarded as trivial properties. Example 7.2(Parallel System). Let us consider a machine with $N$ components, each alternately functioning (up) and defective (down). Suppose that $X_{i}(u)$ denote the states ’ of the $i\mathrm{t}\mathrm{h}$ component, that is, the process with state space $\{0,1\}$ , where $0=‘ \mathrm{u}\mathrm{p}$ and $1=$ ‘down’, and suppose that Assumption 2.1 holds for the family $\{X_{i}(u) : i=1, \ldots , N\}$ . Set $X(u)= \sum_{i=1}^{N}x_{i}(u)$ , and consider $X(u)$ on the $N$-cube. For $X(u)$ , starting at the origin

(state $0$ ), denote by $\overline{\tau}_{0N}$ the first-passage time of $X(u)$ to the opposite vertex (state $N$ ).

If the components are $\mathrm{i}\mathrm{n}_{-}\mathrm{p}\mathrm{a}\mathrm{r}\mathrm{a}\mathrm{l}\mathrm{l}\mathrm{e}1$ , so that the machine functions if and only if at least one component does, then $\tau_{0N}$ is the time to first failure of the machine. The results (iii),

(iv) and $(\underline{\mathrm{v}})$ of Theorem 2.3 show the asymptotc behavior of the mean first-failure time

$\overline{\mu}_{0N}=E[T_{0N}]$ as $Narrow\infty$ : $\overline{\mu}0N\sim 2^{N}$ for large $N$ , if $s=t=1$ . 98

$7.5_{\mathrm{i}}$ Queueing network

Let $X(u)$ be the Markove process in the reduced description with transition probabilities as given in (2.4) in the reduced description. Then $X(u)$ represents a model of a closed queueing network of the Jackson type with two nodes (urn I and urn II), each of which contains $N$ servers whose service times are independent exponentials of parameter $1/N$ , satisfying the following: The total number of customers is $N$ . Each customer, after completion of his service in urn

I (resp. urn II), moves to urn II (resp. urn I) with probability $t$ (resp. $s$ ) and remains in urn I (resp. urn II) with probability $1-t$ (resp. l-s) in order to be served, repeatedly. The results as obtained in the previous section 2 may be extended to the ‘multi-urn’ versions of the classical Ehrenfest model in order that the continuous-time formulation can be applied to closed networks with multi-node. 7.6 Infinite system of queue processes

In the finite system of differential equations (7.3), let $N$ tend to infinity. Then we obtain the (formal) infinite system:

$\frac{d}{du}P_{i,k}^{*}(u)=-sP_{i^{*}k},(u)+sP_{i^{*}+1,k}(u)$ $i,$ , $k=0,1,2,$ $\ldots$ ,

$P_{i,k}^{*}(0)=1$ if $i=k$ , $P_{i,k}^{*}(0)=0$ if $i\neq k$ , which are the differential equations associated with the pure birth process with state space $\{0,1,2, \ldots\}$ . We can obtain another infinite system under suitable scale change. Example 7.3(Infinitely Many Servers). Fix a positive integer $N_{0}$ which is large enough,

$b$ and choose $N$ so that $N>N_{0}$ . Set $s=a/N$ and $t=1-b/N$ . Here $a$ and are positive real numbers, which are chosen so that $0

$N_{0} \frac{d}{du}P_{i,k}(u)$ $=$ $( \frac{N_{0}}{N})i(1-\frac{b}{N})P_{i-1,k}(u)$

$- \{(\frac{N_{0}}{N})i(1-\frac{b}{N})+(\frac{N_{0}}{N})(1-\frac{i}{N})a\}Pi,k(u)$

$+( \frac{N_{0}}{N})(1-\frac{i}{N})$ a $P_{i+1,k}(u)$ .

Let $N$ tend to infinity in the avove equations, assuming that $N_{0}/N\approx 1$ as $Narrow\infty$ . Then, we obtain the following (formal) infinite system:

$N_{0} \frac{d}{du}P_{i}^{*},(ku)$ $=$ $iP_{i-1,k}^{*}(u)-(i+a)P_{i,k}^{*}(u)+aP_{i+1,k}^{*}(u)$ ,

$i,$ $k$ $=$ $0,1,2,$ $\ldots$ ,

$P_{i,k}^{*}(0)$ $=$ 1 if $i=k$ , $P_{i,k}^{*}(0)=0$ if $i\neq k$ , which are the differential equations associated with a birth-death queueing process with infinitely many servers. 99

7.7 Manufacturing system

We consider the stochastic manufacturing system with the surplus level $x(u)\in R^{1}$ and the production rate $y(u)\in R^{1}$ satisfying

$\frac{d}{du}x(u)=y(u)-z$ , $x(0)=x$ ,

$z\in R^{1}$ where is the constant demand rate and $x$ is the initial value of $x(u)$ . Let $S=\{0,1,2, \ldots , N\}$ denote the set of capacity states, and let $K(\epsilon, u)\in S,$ $u_{=}>0$ , denote the total production capacity process of the parallel machines in the system, which is given by a finite state Markov process parameterized by a small number $\epsilon>0$ . Then the production rate $y(u)$ must satisfy $0_{=}<_{y}(u)=$ O. We consider the cost function $J^{\epsilon}(x, k, y(\cdot))$ with $K(\epsilon, \mathrm{O})=k$ defined by

$J^{\epsilon}(X, k, y( \cdot))=E[\int_{0}^{\infty}\exp(-\rho u)G(X(u), y(u))du]$ ,

$G(x, y)$ $x$ where is the running cost of having surplus and production rate $y$ and $\rho>0$ is the $0\leqq y(u)=0=$ discount rate. The problem is to find an admissible control ’ as a function of the past $K(\epsilon, u)$ , that minimizes $J^{\epsilon}(x, k, y(\cdot))$ . According to Sethi [12], we make the following assumptions on $G$ and $K(\epsilon, u)$ : (i) The cost function $G$ satisfies suitable smoothness condition and growth restriction. $0=0=$ ’ is a finite Markov process governed by the infinitesimal generator $Q^{\epsilon}$ of the form

$Q^{\in}=Q^{()(}1+ \frac{1}{\epsilon}Q2)$ , where $Q^{(l)}$ is a $(N+1)\cross(N+1)$ matrix such that $Q^{(l)}=(q_{ij}^{()}l)$ with $q_{ij}^{(l)}=>0$ if $i\neq j$

$q_{ii}^{(l)}=- \sum i\neq jq_{i}^{(l)}j$ $Q^{(2)}$ and ’ for $l=1,2$ . The matrix is weakly irreducible, that is, it has a unique stationary probability distribution. Under these assumtions, Sethi [12] investigates the limiting control problem as $\epsilonarrow 0$ . Sethi [12] also considers $K(\epsilon, u)$ which is modelled by a birth-death process such that the second $Q$ -matrix $Q^{(2)}=(q_{i}^{(2)}j)$ in the infinitesimal generator $Q^{\epsilon}$ has the following expresslon:

$q_{ij}^{(2)}=\lambda_{i}$ if $j=i+1$ , $q_{ij}^{(2)}=\mu_{i}$ if $j=i-1$ ,

$q_{ij}^{(2)}=-(\lambda_{i}+\mu_{i})$ , if $j=i$ , $q_{ij}^{(2)}=0$ otherwise,

$\lambda_{i}$ $Q$ $Q^{(2)}$ with nonnegative constants and $\mu_{i}$ . In the case where the second -matrix is given by the Ehrenfest model (7.2), the value function for the above problem contains several parameters, such as $0<\epsilon\ll 1,0=1$ . We may be concerned with the asymptotic analysis as $\epsilon\downarrow 0,$ $s\uparrow 1,$ $t\uparrow 1$ and $N\uparrow\infty$ . 100

8. Limiting distribution and diffusion approximation

Let $P_{ik}(u)$ be the transition probabilities of the continuous-time Markov chain in the reduced description as given in Theorem 2.2. Then we have the following results for the limiting distribution and diffusion approximation.

8.1 The rate of convergence to the limiting binomial distribution

Theorem 2.2 implies the explicit formula:

$P_{ik}(u)=( \frac{p}{q})^{k}\sum_{x=0}^{N}p^{x}q-x_{I}\mathrm{i}_{i}’(N)x\Lambda’k(x)\exp[-\lambda xu]$ ,

where $\mathrm{A}_{i}’(x)=\mathrm{A}_{i}’(x,p, N)$ and $\lambda=(s+t)/N$ . Using

$K_{n}(0)=1$ , $I\iota_{0}’(X)=1$ , which appear in (ii) of Remark 1.4, we have

$P_{ik}(u)$ $=$ $( \frac{p}{q})^{k}\{q^{N}I\mathrm{t}_{i}^{\nearrow}(0)I\zeta_{k}(0)+\sum_{=x1}p^{x}q^{N-}K_{i(x)\iota_{k}}xI’(X)\exp[-\lambda Xu]\}N$

$=$ $p^{k}qN-k+ \exp[-\lambda u]\overline{\pi}_{k}\sum_{x=1}^{N}p^{x}q^{N-x_{I}}\mathrm{i}’i(x)I\mathrm{i}_{k}’(x)\exp[-\lambda(X-1)u]$ , where

$\overline{\pi}_{k}=(\frac{p}{q})^{k}$ , that is,

$P_{ik}(u)=p^{k}q-k+ON(e^{-})\lambda u$ as $uarrow\infty$ .

In other words the rate of convergence to the limiting (binomial) distribution is expo- nential of order $e^{-\lambda u}$ The constant in the $O$-symbol is easily expressed in terms of the parameters, in fact

$P_{ik}(u)=p^{k}q^{N-k} \{1+\sum_{x=1}^{N}px-qix_{I\mathrm{i}(}\prime x)I\mathrm{i}_{k}’(X)\exp[-\lambda xu]\}$

$=$ $p^{k}q^{Nk}- \{1+Npq^{-1}\frac{1}{pN}$ (Np–i) $\frac{1}{pN}$ (Np–k) $\exp[-\lambda u]+O(\exp[-2\lambda u])\}$

$=$ $p^{k}q^{N-k} \{1+\frac{(Np-i)(Np-k)}{Npq}e^{-\lambda u}+O(e^{-})2\lambda u\}$ ,

since $I\mathrm{i}_{j}’(1)=k_{j1}=k_{1j}=(Np-j)/(pN)$ as given in subsection 7.1. 101

8.2 Convergence to the Ornstein-Uhlenbeck process

The Ehrenfest chain of the model (1.1) is given on the state space $\{0,.1, \ldots, N\}$ with transition probabilities:

$p(i,i-1)= \frac{i}{N}t$ , $p(i, i+1)=(1- \frac{i}{N})s$ , $p(i, i)=1-(1- \frac{i}{N})s-\frac{i}{N}t$ , where $0

$\{-a, -a+1, \ldots , -1,0,1, \ldots, a-1, a\}$ with the following transition probabilities :

$p(k, k-1)$ $=$ $( \frac{k+a}{2a})t=\frac{1}{2}(1+\frac{k}{a})t$ ,

$p(k, k+1)$ $=$ $( \frac{a-k}{2a})s=\frac{1}{2}(1-\frac{k}{a})s$ ,

$p(k, k)$ $=$ $1- \frac{1}{2}(1-\frac{k}{a})s-\frac{1}{2}(1+\frac{k}{a})t$ .

Let $p(n;k, m)$ be the $n$ -step transition probabilities such that $p(n;k, m)=P(X_{n}=m|X_{0}=k)$ . Then it is clear that

$p(n+1;k, m)$

$=$ $p(n;k, m-1)p(m-1, m)+p(n;k, m+1)p(m+1, m)+p(n;k, m)p(m, m)$

$=$ $\frac{1}{2}(1-\frac{m-1}{a})Sp(n;k, m-1)+\frac{1}{2}(1+\frac{m+1}{a})tp(n;k, m+1)$

$+ \{1-\frac{1}{2}(1-\frac{m}{a})S-\frac{1}{2}(1+\frac{m}{a})t\}p(n;k, m)$.

Let us speed up this nonhomogeneous random walk. Let $a(=N/2)arrow\infty$ such that

$\frac{(\triangle x)^{2}}{\triangle u}=\sigma^{2}$ and $a(\triangle u)arrow\nu^{-1}$ with constants $\sigma>0$ and $l\text{ノ}>0$ .

Let us pass from the $n$ -step transition probability to the transition density $f(u, y)=f(x;u, y)$ by taking $n(\triangle u)=u$ , $m(\triangle x)=x$ :

$f(u+\triangle u, y)$ $=$ $\frac{1}{2}(1-\frac{\gamma\gamma l-1}{a})sf(u, y-\triangle y)+\frac{1}{2}(1+\frac{m+1}{a})tf(u, y+\triangle y)$

$+ \{1-\frac{1}{2}(1-\frac{m}{a})s-\frac{1}{2}(1+\frac{m}{a}\mathrm{I}t\}f(u, y)$. 102

Then

$f(u, y)+( \triangle u)\frac{\partial f}{\partial u}+o(\triangle u)$

$=$ $\frac{1}{2}(1-\frac{m-1}{a})s\{f(u, y)-(\triangle y)\frac{\partial f}{\partial y}+\frac{1}{2}(\triangle y)^{2}\frac{\partial^{2}f}{\partial y^{2}}\}$

$+ \frac{1}{2}(1+\frac{m+1}{a})t\{f(u, y)+(\triangle y)\frac{\partial f}{\partial y}+\frac{1}{2}(\triangle y)2_{\frac{\partial^{2}f}{\partial y^{2}}}\}$

$+ \{1-\frac{1}{2}(1-\frac{m}{a})S-\frac{1}{2}(1+\frac{m}{a})t\}f(u, y)+o((\triangle y)^{2})$

$=$ $\{1+\frac{t+s}{2a}\}f(u, y)+\{\frac{t-s}{2}+\frac{m}{2a}(t+s)+\frac{t-s}{2a}\}(\triangle y)\frac{\partial f}{\partial y}$

$+ \{\frac{t+s}{2}+\frac{m}{2a}(t-S)+\frac{t+s}{2a}\}\frac{1}{2}(\triangle y)2\frac{\partial 2f}{\partial y^{2}}+o((\triangle y)^{2})$ .

Observe that $m=y/(\triangle y)$ , so that

$\frac{\partial f}{\partial u}+\frac{o(\triangle u)}{\triangle u}$

$=$ $( \frac{t+s}{2})(\frac{1}{a(\triangle u)})f(u, y)$

$+ \{(\frac{t-s}{2})(\frac{1}{\triangle u})(\triangle y)\frac{\partial f}{\partial y}+(\frac{t+s}{2})(\frac{1}{a(\triangle u)})y\frac{\partial f}{\partial y}$

$+( \frac{t-s}{2})(\frac{1}{a(\triangle u)})(\triangle y)\frac{\partial f}{\partial y}\}$

$+ \{(\frac{t+s}{2})\frac{1}{2}(\frac{(\triangle y)^{2}}{\triangle u})\frac{\partial^{2}f}{\partial y^{2}}+(\frac{t-s}{2})(\frac{1}{a(\triangle u)})y(\triangle y)\frac{\partial^{2}f}{\partial y^{2}}$

$+( \frac{t+s}{2})(\frac{1}{a(\triangle u)})(\triangle y)^{2}\frac{\partial^{2}f}{\partial y^{2}}\}+\frac{o((\triangle y)^{2})}{\triangle u}$ .

$\triangle u$ Assume that $t=s$ and choose and $\triangle y$ so that

$\frac{(\triangle y)^{2}}{\triangle u}arrow\sigma^{2}$ $\frac{1}{a(\triangle u)}arrow\nu$ $(\triangle uarrow \mathrm{O})$ and $(\triangle uarrow \mathrm{O}, \triangle yarrow 0)$ .

Then, we can obtain the following Kolmogorov’s forward equation for the Ornstein- Uhlenbeck process :

$\frac{\partial f}{\partial u}=t\{\nu\frac{\partial}{\partial y}(yf)+\frac{1}{2}\sigma\frac{\partial^{2}f}{\partial y^{2}}2\}$ . 103

References

[1] D. Aldous: Finite-time implications of relaxation times for stochastically monotone processes. Probab. Th. Rel. Fields, 77 (1988) 137-145. [2] R.N. Bhattacharya and E.C. Waymire: Stochastic Processes with Applications (Wiley, New York, 1990). [3] N. H. Bingham: Fluctuation theory for the Ehrenfest urn. Adv. Appl. Prob. 23 (1991) 598-611. [4] H. Dette: On a generalization of the Ehrenfest urn model. J. Appl. Prob. 31 (1994) 930-939. [5] P. and T. Ehrenfest: Uber zwei bekannte Einw\"ande gegen das Boltzmannsche H- Theorem. Physik. Zeit. 8 (1907) 311-314. [6] M. Kac: Random walk and the theory of Brownian motion. Amer. Math. Mon..thly 54 (1947) 369-391. [7] S. Karlin and J. McGregor: Ehrenfest urn models. J. Appl. Prob. 2 (1965) 352-376. [8] M. Kijima: Upper bounds of a measure of dependence and the relaxation time for finite state Markov chains. J. Operations Research Soc. Japan, 32 (1989) 93-102. [9] O. Krafft and M. Schaefer: Mean passage times for tridiagonal transition matrices and a two-parameter Ehrenfest urn model. J. Appl. Prob. 30 (1993) 964-970. [10] J. L. Palacios: Fluctuation theory for the Ehrenfest urn via electric networks. Adv. Appl. Prob. 25 (1993) 472-476. [11] J. L. Palacios: Another look at t,he Ehrenfest urn via electric networks. Adv. Appl. Prob. 26 (1994) 820-824. [12] S.P. Sethi: Some insights into near-optional plans for stochastic manufacturing sys- tems. In G.G. Yin and Q. Zhang $(\mathrm{e}\mathrm{d}_{\mathrm{S}}.)$ : Mathematics of Stochastic Manufacturing $s_{y_{S}}tems_{f}$ Lectures in Appl. Math. 33 (AMS, Providence, 1997), 287-315. [13] L. Tak\’acs: On an urn problem of Paul and Tatiana Ehrenfest. Math. Proc. Camb. Phil. Soc. 86 (1979) 127-130. [14] I. Vincze: \"Uber das Ehrenfestsche Modell der W\"arme\"ubertragung. Arch. Math. 15 (1964) 394-400.

Kiyomasa Narita Department of $\mathrm{I}\mathrm{n}\mathrm{d}\mathrm{u}\mathrm{s}\dot{\mathrm{t}}\mathrm{r}\mathrm{i}\mathrm{a}\mathrm{l}$ Engineering and Management Faculty of Engineering, Kanagawa University Rokkakubashi Kanagawa-ku Yokohama 221-8686, Japan Email: [email protected]